Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Realization of the unconventional photon blockade based on a three-wave mixing system

Open Access Open Access

Abstract

In this paper, the unconventional photon blockade is studied in a three-wave-mixing system with a non-degenerate parametric amplification. A method of only retaining the Fock-state basis in the interference path is used to calculate the optimal analytic conditions of unconventional photon blockade. The numerical results agree well with the analytic conditions, which verifies the validity of this method. Our calculations indicate that the strong photon antibunching can be obtained in the high-frequency mode of the three-wave mixing. And the influence of system parameters on photon blockade is also discussed.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The realization of photon non-classical state [1,2] is a research hotspot of quantum optics and quantum integration technology, which is of great significance to the development and application of optical quantum field in the future. In particular, the development of single photon source technology is the main technical source to promote the development of quantum communication [3], quantum metrology [4] and quantum information technology [5,6], and the single photon source has become the ultimate goal of its product definition based on photon architecture. To this end, the study of the single-photon sources has become a conspicuous focus in quantum physics. The key of single photon source lies in the strong photon antibunching and the sub-Poisson statistics, which requires the system with strong optical nonlinearity to realize the photon blockade (PB) [710].

There are two main physical mechanisms for the realization of PB effect: One is conventional phonon blockade (CPB), which is induced by the splitting of the energy levels. The other is unconventional photon blockade (UPB), which is activated by the quantum destructive interference between different paths. The CPB was first observed in an optical cavity coupled to a single trapped atom [11]. After that, CPB effects were observed in different systems by several experimental groups, including the quantum dots in the photonic crystal system [12] and the quantum electrodynamics system in the circuit cavity [13,14]. At the same time, CPB effect has many different applications [1522]. The physical mechanism of UPB is discovered by Liew and Savona, which dramatically increases the development of this research field [23]. In theory, the UPB effect can be implemented in many systems, such as, optomechanical systems [2427], weak nonlinear photon molecules modes [28], second-order nonlinear systems [22,2932], optical cavities with a quantum dot [33], and bimodal coupled polaritonic cavities [34]. Recently, it has been shown that the two mechanisms are connected [35,36] and they can even arise simultaneously [37,38].

In this paper, we study the UPB in a three-wave-mixing system with a non-degenerate parametric amplification. In order to obtain the optimal analytic conditions of UPB, we truncate the wave function in an effective space by retaining the Fock-state basis in the interference path and neglecting the irrelevant Fock-state basis. The numerical results are calculated by solving the master equation in a truncated Hilbert space, which agree with the optimal analytic conditions. The results indicate that the strong photon antibunching can be obtained in the high-frequency mode of the three wave mixing. Our scheme provides an approach for the UPB in a complex quantum system.

The manuscript is organized as follows: In Sec. 2, we introduce the physical model. In Sec. 3, we illustrate analytical results and physical mechanism of UPB. In Sec. 4, we show the comparison of numerical and analytical solutions for the UPB. Conclusions are given in Sec. 5.

2. Physics model

The three-wave mixing system exists a high-frequency mode $a$ with frequency $\omega _a$ and two-low-frequency modes $b$ and $c$ with frequencies $\omega _b$ and $\omega _c$, respectively. Here, the three-wave mixing mediates the convert a photon with high frequency into two photons with different low frequencies. The Hamiltonian of this system can be written as [39,40]

$$\begin{aligned}\hat{H}_{0}=\omega_{a} \hat{a}^{\dagger} \hat{a}+\omega_{b} \hat{b}^{\dagger} \hat{b}+\omega_{c} \hat{c}^{\dagger} \hat{c}+g\left(\hat{a}^{\dagger} \hat{b} \hat{c}+\hat{c}^{\dagger} \hat{b}^{\dagger} \hat{a}\right) \end{aligned}$$
where $\hat {a}$($\hat {a^{\dagger}}$), $\hat {b}(\hat {b^{\dagger}})$ and $\hat {c}(\hat {c^{\dagger}})$ denotes the annihilation (creation) operator of the three modes, respectively. And $\omega _a$, $\omega _b$ and $\omega _c$ denotes the frequency of the three modes, respectively. $g$ denotes the coefficient of second-order nonlinear interactions, which can be realized with state-of-the-art technology employing the main III-V materials, such as GaAs, GaN, and AIN [20]. In order to guarantee the maximum value of $g$, we need to satisfy the energy conservation condition $\omega _a=\omega _b+\omega _c$.

The Hamiltonian with non-degenerate parametric amplification and weak driving field can be expressed as:

$$\begin{aligned} \hat{H}=&\omega_a\hat{a}^{\dagger}\hat{a}+\omega_b\hat{b}^{\dagger}\hat{b}+\omega_c\hat{c}^{\dagger}\hat{c}+g(\hat{a}^{\dagger}\hat{b}\hat{c}+\hat{c}^{\dagger}\hat{b}^{\dagger}\hat{a})\\ &+F_a(\hat{a}^{\dagger}e^{{-}i\omega_{la}t}+\hat{a}e^{i\omega_{la}t})\\ &+E(\hat{b}\hat{c}e^{{-}i(\omega_{lb}+\omega_{lc})t}+\hat{b}^{\dagger}\hat{c}^{\dagger}e^{i(\omega_{lb}+\omega_{lc})t}), \end{aligned}$$

In order to achieve photon blockade, weak driving is necessary, which is similar to the laser controlled Rydberg atomic process [4145]. Here $F_a$ represents the weak drive intensity of mode $a$ with the frequency $\omega _{la}$, and $E$ represents the non-degenerate parametric amplification coefficient [4649] of mode $b$ and $c$ with frequencies $\omega _{lb}$ and $\omega _{lc}$.

For convenience, we would like to turn to a rotation framework subject to the operator $\hat {U}(t)=e^{i t(\omega _{la}\hat {a}^{\dagger }\hat {a}+\omega _{lb}\hat {b}^{\dagger }\hat {b}+\omega _{lc}\hat {c}^{\dagger }\hat {c})}$, which leads to an effective Hamiltonian $\hat {H}_{eff}=\hat {U}\hat {H}\hat {U}^{\dagger }-i \hat {U}dU^{\dagger }/dt$ as

$$\begin{aligned} \hat{H}_{eff}=&\Delta_a\hat{a}^{\dagger}\hat{a}\ +\Delta_b\hat{b}^{\dagger}\hat{b}+\Delta_c\hat{c}^{\dagger}\hat{c}+g(\hat{a}^\dagger\hat{b}\hat{c}+\hat{a}\hat{b}^{\dagger}\hat{c}^{\dagger})\\ &+F_a(\hat{a}^{\dagger}+\hat{a})+E(\hat{b}\hat{c}+\hat{b}^{\dagger}\hat{c}^{\dagger}), \end{aligned}$$
where $\Delta _a=\omega _a-\omega _{la}$, $\Delta _b=\omega _b-\omega _{lb}$ and $\Delta _c=\omega _c-\omega _{lc}$ represent the detunings, which satisfy the relationship $\Delta _a=\Delta _b+\Delta _c$.

3. Analytical results

The Fock-state basis of the system is denoted by $|m,n,p\rangle$ with the number $m$, $n$ and $p$ denoting the photon number in mode $a$, $b$ and $c$, respectively. The system is truncated to two-photon excitation. And the Fock-state basis is shown in Fig. 1(a), where we consider the relation $\omega _b=2\omega _c$. To obtain the optimal analytic conditions for UPB, we analyse the interference paths shown in Fig. 1(b). There are so many Fock-state basis in Fig. 1(a), which lead to that it is difficult to calculate the optimal conditions for UPB. Here we only keep the Fock-state basis in Fig. 1(b), which are related to the interference paths. In this case, the the wave function of the system can be expressed as:

$$\begin{aligned} |\psi\rangle =& C_{000}|000\rangle+C_{100}|100\rangle+C_{200}|200\rangle+C_{111}|111\rangle\\ &+C_{022}|022\rangle. \end{aligned}$$

Considering the impact of the environment involved in the loss of the modes. We introduce the non-Hermitian Hamiltonian, which can be writen as

$$\tilde{H}=\hat{H}_{eff}-i\frac{\kappa}{2}\hat{a}^\dagger\hat{a} -i\frac{\kappa}{2}\hat{b}^\dagger\hat{b}-i\frac{\kappa}{2}\hat{c}^\dagger\hat{c}$$

 figure: Fig. 1.

Fig. 1. (a) The Fock-state basis of the system. (b) The Fock-state basis, which are involved in the interference paths.

Download Full Size | PDF

Substituting the wave function Eq. (4) and non-Hermitian Hamiltonian Eq. (5) into the Schrödinger equation $i\partial _t|\psi \rangle =\widetilde {H}|\psi \rangle$, we can obtain the coupling equations for the coefficients,

$$\begin{aligned} i\dot{C}_{000}&=F_a C_{100}+E C_{011}\\ i\dot{C}_{100}&=F_a C_{000}+(\Delta_a-\frac{i\kappa}{2})C_{100}+g C_{011}+F_aC_{200},\\ i\dot{C}_{011}&=E C_{000}+g C_{100}+(\Delta_a-2\frac{i\kappa}{2})C_{011}+F_aC_{111}+2EC_{022},\\ i\dot{C}_{200}&=\sqrt{2}F_a C_{100}+2(\Delta_a-\frac{i\kappa}{2})C_{200}+\sqrt{2}gC_{111},\\ i\dot{C}_{111}&=E C_{100}+ F_aC_{011}+\sqrt{2}g C_{200}+(2\Delta_a-3\frac{i\kappa}{2})C_{111}+2gC_{022}\\ i\dot{C}_{022}&=2E C_{011}+2g C_{111}+2(\Delta_a-2\frac{i\kappa}{2})C_{022} \end{aligned}$$

By solving the coupling equation of the coefficients, we consider the steady state solution of this system

$$\begin{aligned} &F_a C_{100}+E C_{011}=0,\\ &F_a C_{000}+(\Delta_a-\frac{i\kappa}{2})C_{100}+g C_{011}+F_aC_{200}=0,\\ &E C_{000}+g C_{100}+(\Delta_a-2\frac{i\kappa}{2})C_{011}+F_aC_{111}+2EC_{022}=0,\\ &\sqrt{2}F_a C_{100}+2(\Delta_a-\frac{i\kappa}{2})C_{200}+\sqrt{2}gC_{111}=0,\\ &E C_{100}+ F_aC_{011}+\sqrt{2}g C_{200}+(2\Delta_a-3\frac{i\kappa}{2})C_{111}+2gC_{022}=0,\\ &2E C_{011}+2g C_{111}+2(\Delta_a-2\frac{i\kappa}{2})C_{022}=0, \end{aligned}$$
where $\Delta _a=\Delta _b+\Delta _c$. To implement UPB, the driver condition must satisfy $F_a \ll \kappa$ and $E \ll \kappa$, and the first equation in Eq. (7) is approximatively satisfied. The Eq. (7) can be solved and we set $|C_{200}|=0$, the optimal analytic conditions of UPB are obtained as
$$\begin{aligned} &[ \quad g_1=\frac{\lambda}{\sqrt{2}}, \quad E_1={-}\frac{1}{\varphi}\varepsilon \quad],\\ &[ \quad g_2={-}\frac{\lambda}{\sqrt{2}}, \quad E_2=\frac{1}{\varphi}\varepsilon \quad], \end{aligned}$$
where the parameters $\lambda =\sqrt {68\kappa ^2+\frac {18\kappa ^4}{\Delta ^2}+64\Delta ^2-\frac {\sqrt {(6\kappa ^2+13\Delta ^2)^2(9\kappa ^2+3\kappa ^2\Delta ^2+23\Delta ^4)}}{\Delta ^2}}$ , $\varphi =-648\kappa ^8-3882\kappa ^6\Delta ^2-7121\kappa ^4\Delta ^4-2542\kappa ^2\Delta ^6+2717\Delta ^8$ and $\varepsilon =356F\kappa ^6\Delta \lambda -1987\kappa ^4\Delta ^3\lambda -3606F\kappa ^2\Delta ^5\lambda -2115F\Delta ^7\lambda +8F\kappa ^2\Delta ^3\lambda ^3+18F\Delta ^5\lambda ^3$. When the optimal conditions Eqs. (8) are satisfied, the UPB will occur in high frequency mode $a$, which can be verified by the numerical simulations.

4. Comparison of the numerical results and analytic results

4.1 Numerical solution

The dynamics of the density matrix $\hat {\rho }$ of the system is governed by

$$\begin{aligned} \frac{\partial\hat{\rho}}{\partial t}=&-i[\hat{H}_{eff},\rho]+\frac{\kappa_a}{2}(\bar{n}_{th}+1)(2\hat{a}\hat{\rho}\hat{a}^\dagger+\frac{1}{2}\hat{a}^\dagger\hat{a}\hat{\rho}+\frac{1}{2}\hat{\rho}\hat{a}^\dagger\hat{a})\\ &+\frac{\kappa_b}{2}(\bar{n}_{th}+1)(2\hat{b}\hat{\rho}\hat{b}^\dagger+\frac{1}{2}\hat{b}^\dagger\hat{b}\hat{\rho}+\frac{1}{2}\hat{\rho}\hat{b}^\dagger\hat{b})\\ &+\frac{\kappa_c}{2}(\bar{n}_{th}+1)(2\hat{c}\hat{\rho}\hat{c}^\dagger+\frac{1}{2}\hat{c}^\dagger\hat{c}\hat{\rho}+\frac{1}{2}\hat{\rho}\hat{c}^\dagger\hat{c})\\ &+\frac{\kappa_a}{2}\bar{n}_{th}(2\hat{a}^\dagger\hat{\rho}\hat{a}+\frac{1}{2}\hat{a}\hat{a}^\dagger\hat{\rho}+\frac{1}{2}\hat{\rho}\hat{a}\hat{a}^\dagger)\\ &+\frac{\kappa_b}{2}\bar{n}_{th}(2\hat{b}^\dagger\hat{\rho}\hat{b}+\frac{1}{2}\hat{b}\hat{b}^\dagger\hat{\rho}+\frac{1}{2}\hat{\rho}\hat{b}\hat{b}^\dagger)\\ &+\frac{\kappa_c}{2}\bar{n}_{th}(2\hat{c}^\dagger\hat{\rho}\hat{c}+\frac{1}{2}\hat{c}\hat{c}^\dagger\hat{\rho}+\frac{1}{2}\hat{\rho}\hat{c}\hat{c}^\dagger), \end{aligned}$$
where $\kappa _a$, $\kappa _b$ and $\kappa _c$ denote the decay rates of mode $a$, mode $b$ and mode $c$, respectively. Without loss of generality, the decay rates of the cavity modes are assumed to be equal, i.e., $\kappa _a=\kappa _b=\kappa _c=\kappa$. $\bar {n}_{th}$=$\{\exp {[\hbar \omega /(\kappa _BT)-1]}\}^{-1}$ is the mean number of thermal photons, $\kappa _B$ is the Boltzmann constant, and $T$ is the reservoir temperature at thermal equilibrium.

In this paper, we only concern the zero-delay-time second-order correlation function in the steady state, so we need the steady-state density operator $\hat {\rho }_s$, which can be obtained by setting $\partial \hat {\rho }/\partial t=0$. The zero-delay-time second-order correlation function for mode $a$ is defined by

$$g^{(2)}(0)=\frac{\langle \hat{a}^{\dagger}\hat{a}^{\dagger}\hat{a}~\hat{a}\rangle} {\langle \hat{a}^{\dagger}\hat{a}\rangle^2},$$

The second-order correlation function $g^{(2)}(0)<1$ corresponds to the sub-Poissonian statistics.

4.2 Comparison of numerical solutions and analytical conditions

In the following, the numerical results are provided to study the UPB effect by plotting the second-order correlation function $g^{(2)}(0)$ versus the system parameters. The Hilbert spaces of the system are truncated to five dimensions for cavity modes $a$, $b$ and $c$ respectively. And for convenience, we reset all parameters to units of dissipation rate $\kappa$.

First, we numerically study the UPB under the zero temperature ($\bar {n}_{th}=0$) and the results are compared with the optimal analytic condition shown in Eq. (8). In Fig. 2(a), we logarithmically plot $g^{(2)}(0)$ as a function of $g_1/\kappa$ and $\Delta _a/\kappa$ for mode $a$, and the other parameters are $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, and $F_{a}/\kappa =0.01$. According to the analytical results, we set the $E_1/\kappa =-\frac {1}{\varphi }\varepsilon$, where the parameters $\varphi =-648\kappa ^8-3882\kappa ^6\Delta ^2-7121\kappa ^4\Delta ^4-2542\kappa ^2\Delta ^6+2717\Delta ^8$ and $\varepsilon =356F\kappa ^6\Delta \lambda -1987\kappa ^4\Delta ^3\lambda -3606F\kappa ^2\Delta ^5\lambda -2115F\Delta ^7\lambda +8F\kappa ^2\Delta ^3\lambda ^3+18F\Delta ^5\lambda ^3$. The numerical results show that, the UPB can occur in this system. The valleys in $g^{(2)}(0)< 1$ correspond to the strong photon antibunching. The analytic condition is denoted by the white dotted line, which agrees well with numerical simulations. In Fig. 2(b), we logarithmically plot $g^{(2)}(0)$ as a function of $g_2/\kappa$ and $\Delta _a/\kappa$ for mode $a$, and the other parameters are $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.01$, and $E1/\kappa =\frac {1}{\varphi }\varepsilon$, which is consistent with the results in Fig. 2(a), the valleys in $g^{(2)}(0)< 1$ correspond to the strong photon antibunching and the analytic solution is denoted by the white dotted line, the analytical solutions and numerical simulations agree well in this situation. According to the results of Fig. 2(a) and Fig. 2(b), it is not difficult to find that the blockade regions is symmetrically distributed, which is in complete agreement with the results obtained by analytical calculation shown in Eqs. (8). Therefore, the method of preserving effective interference path adopted by us in dealing with this model is effective. In Fig. 2(c), we logarithmically plot $g^{(2)}(0)$ as a function of $E_1/\kappa$ and $\Delta _a/\kappa$ for mode $a$, where the shared parameters $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, and $F_{a}/\kappa =0.01$. According to the analytical results, we set the $g_1/\kappa =\frac {\lambda }{\sqrt {2}}$, and the parameters $\lambda =\sqrt {68\kappa ^2+\frac {18\kappa ^4}{\Delta ^2}+64\Delta ^2-\frac {\sqrt {(6\kappa ^2+13\Delta ^2)^2(9\kappa ^2+3\kappa ^2\Delta ^2+23\Delta ^4)}}{\Delta ^2}}$. Based on the results shown in the figure, the UPB can occur in this system, where the valleys in $g^{(2)}(0)< 1$ correspond to the strong photon antibunching, the shape of which is similar to the hyperbolic curve, and the white dotted lines represent analytic conditions. In Fig. 2(d), we logarithmically plot $g^{(2)}(0)$ as a function of $E2/\kappa$ and $\Delta _a/\kappa$ for mode $a$, where the shared parameters $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.01$ and $g2/\kappa =-\frac {\lambda }{\sqrt {2}}$. The valleys in $g^{(2)}(0)< 1$ correspond to the strong photon antibunching and the analytic solution is denoted by the white dotted line.In order to further compare and analyze the numerical and analytical results, we discuss the various coefficients appearing in the system model, and then determine the optimal coefficients setting for the UPB. In Fig. 3(a), we logarithmically plot $g^{(2)}(0)$ as a function of $\Delta _a/\kappa$, under different detuning $\Delta _a/\kappa$, where $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.1$, and $E/\kappa =\pm \frac {1}{\varphi }\varepsilon$. In the blue solid line we set the detuning $\Delta _a/\kappa =2$, in the black dotted line $\Delta _a/\kappa =3$, and in the red point line $\Delta _a/\kappa =4$. UPB occurs in all three curves, and each curve has two strong blockade valleys, which are basically match with the optimal analytic conditions. With the increasing of the parameter $\Delta _a/\kappa$, the intensity of UPB increases gradually. At the same time, the absolute value of the second-order nonlinear coefficient $g/\kappa$ also increase. Therefore, we cannot determine that a larger $\Delta _a/\kappa$ value is favorable for this system to implement UPB. Next, we choose $\Delta _a/\kappa =3$ for further analysis of the other coefficients in the system.

 figure: Fig. 2.

Fig. 2. (a) Logarithmic plot of zero-delay-time second-order correlation functions $g^{(2)}(0)$ as a function of $g_1/\kappa$ and $\Delta _a/\kappa$ for mode $a$, where the shared parameters $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.01$ and $E_1/\kappa =-\frac {1}{\varphi }\varepsilon$. (b) Logarithmic plot of $g^{(2)}(0)$ as a function of $g_2/\kappa$ and $\Delta _a/\kappa$ for mode $a$, where the shared parameters $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.01$ and $E_2/\kappa =\frac {1}{\varphi }\varepsilon$. (c) Logarithmic plot of $g^{(2)}(0)$ as a function of $E1/\kappa$ and $\Delta _a/\kappa$ for $\Delta _b=\Delta _a/4$, $\Delta _c=3\Delta _a/4$, $F_{a}/\kappa =0.01$ and $g_1/\kappa =-\frac {\lambda }{\sqrt {2}}$. (d) Logarithmic plot of $g^{(2)}(0)$ as a function of $E2/\kappa$ and $\Delta _a/\kappa$ for $\Delta _b=\Delta _a/4$, $\Delta _c=3\Delta _a/4$, $F_{a}/\kappa =0.01$ and $g_2/\kappa =\frac {\lambda }{\sqrt {2}}$.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. (a) Under different detunings $\Delta _a/\kappa$, we logarithmically plot of $g^{(2)}(0)$ vs the second-order nonlinear interaction strength $g/\kappa$, with $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.1$ and $E/\kappa =\pm \frac {1}{\varphi }\varepsilon$ (b) Under different weak drive factor $F_{a}/\kappa$, we logarithmically plot of $g^{(2)}(0)$ vs the nondegenerate parametric amplification factor. $E/\kappa$, with $\Delta _a/\kappa =3$, $\Delta _b=\Delta _a/8$, $\Delta _c=7\Delta _a/8$ and $g/\kappa =\pm \frac {\lambda }{\sqrt {2}}$.

Download Full Size | PDF

In Fig. 3(b), we logarithmically plot $g^{(2)}(0)$ as a function of $E/\kappa$ under different $F_a/\kappa$, where $\Delta _a/\kappa =3$, $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$ and $g/\kappa =\pm \frac {\lambda }{\sqrt {2}}$. In the blue solid line we set the detuning $F_a/\kappa =0.05$, in the black dotted line $F_a/\kappa =0.08$ and in the red point line $F_a/\kappa =0.1$. The UPB effect occurs on all curves, and each curve has two dips, which corresponds to the analytic conditions. With the increase of the driving strength $F_a/\kappa$, the UPB effect of the system became weaker and weaker. Meanwhile, the absolute value of the non-degenerate parametric amplification coefficient $E/\kappa$ also moves in the direction of the increase.

Finally, we investigate the UPB effect under nonzero temperatures. In Fig. 4, we plot $g^{(2)}(0)$ vs the second-order nonlinear interaction strength $g/\kappa$ with $\Delta _a/\kappa =3$, $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.05$, and $E/\kappa =0.05$. In the black solid line we set $\bar {n}_{th}=0$, in the red dotted line $\bar {n}_{th}=0.01$, and in the blue point line $\bar {n}_{th}=0.02$. The results show that the strongest UPB point appears at $g/\kappa =-2.32$, just as predicted by analytical conditions. Moreover, when $\bar {n}_{th}$ changes in a large ranges, the UPB does not change significantly, which indicates that this scheme is not sensitive to the change of the reservoir temperature.

 figure: Fig. 4.

Fig. 4. Under different thermal photons $\bar {n}_{th}$, we logarithmically plot of the $g^{(2)}(0)$ vs he second-order nonlinear interaction strength $g/\kappa$, with $\Delta _a/\kappa =3$, $\Delta _b=\Delta _a/5$, $\Delta _c=4\Delta _a/5$, $F_{a}/\kappa =0.05$ and $E/\kappa =0.05$.

Download Full Size | PDF

5. Conclusions

In summary, we have investigated the unconventional photon blockade in a three-wave-mixing system, where the three-wave-mixing mediates can convert a photon with high frequency into two photons with different low frequencies. We simplify the complex wave function composition of the system by preserving the state on the effective interference paths. By analytic calculation, we obtain the optimal condition for strong antibunching, and the discussions of the optimal conditions are presented. Through numerical calculation, we find that the system can realize UPB in the high-frequency mode. According to comparisons of analytical conditions and numerical calculations, ther are in good agreement. The results show that the method of preserving effective interference paths is effective in dealing with this complex model to achieve the UPB effect. So, this method may be useful for dealing with other complex models, and we will discuss it further in the following research. The coefficients in the system are discussed in detail, and the UPB immunes to reservoir temperature.

Funding

National Natural Science Foundation of China (11965017, 11647054); Education Department of Jiangxi Province (GJJ180873); Department of Science and Technology of Jilin Province (2018-0520165JH); Natural Science Foundation of Jilin Province (JJKH20181088KJ).

Disclosures

The authors declare no conflicts of interest.

References

1. L. Davidovich, “Sub-Poissonian processes in quantum optics,” Rev. Mod. Phys. 68(1), 127–173 (1996). [CrossRef]  

2. A. J. Shields, “Semiconductor quantum light sources,” Nat. Photonics 1(4), 215–223 (2007). [CrossRef]  

3. V. Scarani, H. Bechmann-Pasquinucci, N. J. Cerf, M. Dusek, N. Lǔtkenhaus, and M. Peev, “The security of practical quantum key distribution,” Rev. Mod. Phys. 81(3), 1301–1350 (2009). [CrossRef]  

4. V. Giovannetti, S. Lloyd, and L. Maccone, “Advances in quantum metrology,” Nat. Photonics 5(4), 222–229 (2011). [CrossRef]  

5. E. Knill, R. Laflamme, and G. J. Milburn, “A scheme for efficient quantum computation with linear optics,” Nature 409(6816), 46–52 (2001). [CrossRef]  

6. P. Kok, W. J. Munro, K. Nemoto, T. C. Ralph, J. P. Dowling, and G. J. Milburn, “Linear optical quantum computing with photonic qubits,” Rev. Mod. Phys. 79(1), 135–174 (2007). [CrossRef]  

7. A. ImamŌglu, H. Schmidt, G. Woods, and M. Deutsch, “Strongly Interacting Photons in a Nonlinear Cavity,” Phys. Rev. Lett. 79(8), 1467–1470 (1997). [CrossRef]  

8. H. Z. Shen, Y. H. Zhou, H. D. Liu, G. C. Wang, and X. Yi, “Exact optimal control of photon blockade with weakly nonlinear coupled cavities,” Opt. Express 23(25), 32835–32858 (2015). [CrossRef]  

9. H. Z. Shen, S. Xu, Y. H. Zhou, G. Wang, and X. X. Yi, “Unconventional photon blockade from bimodal driving and dissipations in coupled semiconductor microcavities,” J. Phys. B 51(3), 035503 (2018). [CrossRef]  

10. H. Z. Shen, C. Shang, Y. H. Zhou, and X. X. Yi, “Controllable dissipation of a qubit coupled to an engineering reservoir,” Phys. Rev. A 98(2), 023856 (2018). [CrossRef]  

11. K. M. Birnbaum, A. Boca, R. Miller, A. D. Boozer, T. E. Northup, and H. J. Kimble, “Photon blockade in an optical cavity with one trapped atom,” Nature 436(7047), 87–90 (2005). [CrossRef]  

12. A. Faraon, I. Fushman, D. Englund, N. Stoltz, P. Petroff, and J. Vuckovic, “Coherent generation of non-classical light on a chip via photon-induced tunnelling and blockade,” Nat. Phys. 4(11), 859–863 (2008). [CrossRef]  

13. A. J. Hoffman, S. J. Srinivasan, S. Schmidt, L. Spietz, J. Aumentado, H. E. Tureci, and A. A.Houck, “Dispersive Photon Blockade in a Superconducting Circuit,” Phys. Rev. Lett. 107(5), 053602 (2011). [CrossRef]  

14. Y. X. Liu, X. W. Xu, A. Miranowicz, and F. Nori, “From blockade to transparency: controllable photon transmission through a circuit QED system,” Phys. Rev. A 89(4), 043818 (2014). [CrossRef]  

15. L. Tian and H. J. Carmichael, “Quantum trajectory simulations of two-state behavior in an optical cavity containing one atom,” Phys. Rev. A 46(11), R6801–R6804 (1992). [CrossRef]  

16. M. J. Werner and A. Imamoglu, “Photon-photon interactions in cavity electromagnetically induced transparency,” Phys. Rev. A 61(1), 011801 (1999). [CrossRef]  

17. R. J. Brecha, P. R. Rice, and M. Xiao, “two-level atoms in a driven optical cavity: Quantum dynamics of forward photon scattering for weak incident fields,” Phys. Rev. A 59(3), 2392–2417 (1999). [CrossRef]  

18. P. Rabl, “Photon Blockade Effect in Optomechanical Systems,” Phys. Rev. Lett. 107(6), 063601 (2011). [CrossRef]  

19. A. Nunnenkamp, K. Børkje, and S. M. Girvin, “Single-Photon Optomechanics,” Phys. Rev. Lett. 107(6), 063602 (2011). [CrossRef]  

20. A. Majumdar and D. Gerace, “Single-photon blockade in doubly resonant nanocavities with second-order nonlinearity,” Phys. Rev. B 87(23), 235319 (2013). [CrossRef]  

21. H. Z. Shen, Y. H. Zhou, and X. X. Yi, “Quantum optical diode with semiconductor microcavities,” Phys. Rev. A 90(2), 023849 (2014). [CrossRef]  

22. H. Y. Lin, X. Q. Wang, Z. H. Yao, and D. D. Zou, “Kerr-nonlinearity enhanced conventional photon blockade in a second-order nonlinear system,” Opt. Express 28(12), 17643–17652 (2020). [CrossRef]  

23. T. C. H. Liew and V. Savona, “Single Photons from Coupled Quantum Modes,” Phys. Rev. Lett. 104(18), 183601 (2010). [CrossRef]  

24. H. J. Carmichael, “Two-Level Atoms and Spontaneous Emission,” Phys. Rev. Lett. 55(25), 2790–2793 (1985). [CrossRef]  

25. M. Bamba, A. Imamoğlu, I. Carusotto, and C. Ciuti, “Origin of strong photon antibunching in weakly nonlinear photonic molecules,” Phys. Rev. A 83(2), 021802 (2011). [CrossRef]  

26. X. W. Xu and Y. J. Li, “Antibunching photons in a cavity coupled to an optomechanical system,” J. Phys. B: At., Mol. Opt. Phys. 46(3), 035502–035508 (2013). [CrossRef]  

27. V. Savona, “Unconventional photon blockade in coupled optomechanical systems,” arXiv 1302.5937v2.

28. X. W. Xu and Y. Li, “Strongly correlated two-photon transport in a one-dimensional waveguide coupled to a weakly nonlinear cavity,” Phys. Rev. A 90(3), 033809 (2014). [CrossRef]  

29. D. Gerace and V. Savona, “Unconventional photon blockade in doubly resonant microcavities with second-order nonlinearity,” Phys. Rev. A 89(3), 031803 (2014). [CrossRef]  

30. Y. H. Zhou, H. Z. Shen, X. Q. Shao, and X. X. Yi, “Strong photon antibunching with weak second-order nonlinearity under dissipation and coherent driving,” Opt. Express 24(15), 17332–17344 (2016). [CrossRef]  

31. Y. H. Zhou, H. Z. Shen, X. Y. Zhang, and X. X. Yi, “Zero eigenvalues of a photon blockade induced by a non-Hermitian Hamiltonian with a gain cavity,” Phys. Rev. A 97(4), 043819 (2018). [CrossRef]  

32. Y. H. Zhou, H. Z. Shen, and X. X. Yi, “Unconventional photon blockade with second-order nonlinearity,” Phys. Rev. A 92(2), 023838 (2015). [CrossRef]  

33. A. Majumdar, M. Bajcsy, A. Rundquist, and J. Vǔckovic, “Loss-Enabled Sub-Poissonian Light Generation in a Bimodal Nanocavity,” Phys. Rev. Lett. 108(18), 183601 (2012). [CrossRef]  

34. M. Bamba and C. Ciuti, “Counter-polarized single-photon generation from the auxiliary cavity of a weakly nonlinear photonic molecule,” Appl. Phys. Lett. 99(17), 171111 (2011). [CrossRef]  

35. E. Z. Casalengua, J. C. L. Carreo, F. P. Laussy, and E. D. Valle, “Photon Statistics: Conventional and Unconventional Photon Statistics,” Laser Photonics Rev. 14(6), 2070032 (2020). [CrossRef]  

36. M. A. Lemonde, N. Didier, and A. A. Clerk, “Antibunching and unconventional photon blockade with Gaussian squeezed states,” Phys. Rev. A 90(6), 063824 (2014). [CrossRef]  

37. M. Bamba and C. Ciuti, “Nonlinear Quantum Optics with Trion Polaritons in 2D Monolayers: Conventional and Unconventional Photon Blockade,” Phys. Rev. Lett. 125(19), 197402 (2020). [CrossRef]  

38. C. Noh, “Emission of single photons in the weak coupling regime of the Jaynes Cummings model,” Sci. Rep. 10(1), 16076 (2020). [CrossRef]  

39. X. Guo, C. L. Zou, H. Jung, and H. X. Tang, “On-Chip Strong Coupling and Efficient Frequency Conversion between Telecom and Visible Optical Modes,” Phys. Rev. Lett. 117(12), 123902 (2016). [CrossRef]  

40. Y. H. Zhou, H. Z. Shen, X. Y. Luo, Y. Wang, F. Gao, and C. Y. Xin, “Tunable three-wave-mixing-induced transparency,” Phys. Rev. A 96(6), 063815 (2017). [CrossRef]  

41. R. H. Zheng, Y. H. Kang, S.-L. Su, J. Song, and Y. Xia, “Robust and high-fidelity nondestructive Rydberg parity meter,” Phys. Rev. A 102(1), 012609 (2020). [CrossRef]  

42. C. Y. Guo, L.-L. Yan, S. Zhang, S.-L. Su, and W. B. Li, “Optimized geometric quantum computation with a mesoscopic ensemble of Rydberg atoms,” Phys. Rev. A 102(4), 042607 (2020). [CrossRef]  

43. R. Li, D. M. Yu, S.-L. Su, and J. Qian, “Periodically driven facilitated high-efficiency dissipative entanglement with Rydberg atoms,” Phys. Rev. A 101(4), 042328 (2020). [CrossRef]  

44. S.-L. Su, F.-Q. Guo, L. Tian, X.-Y. Zhu, L.-L. Yan, E.-J. Liang, and M. Feng, “Nondestructive Rydberg parity meter and its applications,” Phys. Rev. A 101(1), 012347 (2020). [CrossRef]  

45. C. P. Shen, J.-L Wu, S.-L. Su, and E. J. Liang, “Construction of robust Rydberg controlled-phase gates,” Opt. Lett. 44(8), 2036–2039 (2019). [CrossRef]  

46. C. A. Holmes, “Photon-number-state preparation in nondegenerate parametric amplification,” Phys. Rev. A 39(5), 2493–2501 (1989). [CrossRef]  

47. G. Björk and Y. Yamamoto, “Generation and amplification of numbe states by nondegenerate parametric oscillators with idler-measurement feedback,” Phys. Rev. A 37(1), 125–147 (1988). [CrossRef]  

48. M. S. Abdalla, F. A. A. El-Orany, and J. Peřina, “Quantum statistical properties of nondegenerate optical parametric symmetric coupler,” J. Phys. A: Math. Gen. 32(19), 3457–3483 (1999). [CrossRef]  

49. R. Y. Teh, F.-X. Sun, R. E. S. Polkinghorne, Q. Y. He, Q. Gong, P. D. Drummond, and M. D. Reid, “Dynamics of transient cat states in degenerate parametric oscillation with and without nonlinear Kerr interactions,” Phys. Rev. A 101(4), 043807 (2020). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. (a) The Fock-state basis of the system. (b) The Fock-state basis, which are involved in the interference paths.
Fig. 2.
Fig. 2. (a) Logarithmic plot of zero-delay-time second-order correlation functions $g^{(2)}(0)$ as a function of $g_1/\kappa$ and $\Delta _a/\kappa$ for mode $a$ , where the shared parameters $\Delta _b=\Delta _a/5$ , $\Delta _c=4\Delta _a/5$ , $F_{a}/\kappa =0.01$ and $E_1/\kappa =-\frac {1}{\varphi }\varepsilon$ . (b) Logarithmic plot of $g^{(2)}(0)$ as a function of $g_2/\kappa$ and $\Delta _a/\kappa$ for mode $a$ , where the shared parameters $\Delta _b=\Delta _a/5$ , $\Delta _c=4\Delta _a/5$ , $F_{a}/\kappa =0.01$ and $E_2/\kappa =\frac {1}{\varphi }\varepsilon$ . (c) Logarithmic plot of $g^{(2)}(0)$ as a function of $E1/\kappa$ and $\Delta _a/\kappa$ for $\Delta _b=\Delta _a/4$ , $\Delta _c=3\Delta _a/4$ , $F_{a}/\kappa =0.01$ and $g_1/\kappa =-\frac {\lambda }{\sqrt {2}}$ . (d) Logarithmic plot of $g^{(2)}(0)$ as a function of $E2/\kappa$ and $\Delta _a/\kappa$ for $\Delta _b=\Delta _a/4$ , $\Delta _c=3\Delta _a/4$ , $F_{a}/\kappa =0.01$ and $g_2/\kappa =\frac {\lambda }{\sqrt {2}}$ .
Fig. 3.
Fig. 3. (a) Under different detunings $\Delta _a/\kappa$ , we logarithmically plot of $g^{(2)}(0)$ vs the second-order nonlinear interaction strength $g/\kappa$ , with $\Delta _b=\Delta _a/5$ , $\Delta _c=4\Delta _a/5$ , $F_{a}/\kappa =0.1$ and $E/\kappa =\pm \frac {1}{\varphi }\varepsilon$ (b) Under different weak drive factor $F_{a}/\kappa$ , we logarithmically plot of $g^{(2)}(0)$ vs the nondegenerate parametric amplification factor. $E/\kappa$ , with $\Delta _a/\kappa =3$ , $\Delta _b=\Delta _a/8$ , $\Delta _c=7\Delta _a/8$ and $g/\kappa =\pm \frac {\lambda }{\sqrt {2}}$ .
Fig. 4.
Fig. 4. Under different thermal photons $\bar {n}_{th}$ , we logarithmically plot of the $g^{(2)}(0)$ vs he second-order nonlinear interaction strength $g/\kappa$ , with $\Delta _a/\kappa =3$ , $\Delta _b=\Delta _a/5$ , $\Delta _c=4\Delta _a/5$ , $F_{a}/\kappa =0.05$ and $E/\kappa =0.05$ .

Equations (9)

Equations on this page are rendered with MathJax. Learn more.

H ^ = ω a a ^ a ^ + ω b b ^ b ^ + ω c c ^ c ^ + g ( a ^ b ^ c ^ + c ^ b ^ a ^ ) + F a ( a ^ e i ω l a t + a ^ e i ω l a t ) + E ( b ^ c ^ e i ( ω l b + ω l c ) t + b ^ c ^ e i ( ω l b + ω l c ) t ) ,
H ^ e f f = Δ a a ^ a ^   + Δ b b ^ b ^ + Δ c c ^ c ^ + g ( a ^ b ^ c ^ + a ^ b ^ c ^ ) + F a ( a ^ + a ^ ) + E ( b ^ c ^ + b ^ c ^ ) ,
| ψ = C 000 | 000 + C 100 | 100 + C 200 | 200 + C 111 | 111 + C 022 | 022 .
H ~ = H ^ e f f i κ 2 a ^ a ^ i κ 2 b ^ b ^ i κ 2 c ^ c ^
i C ˙ 000 = F a C 100 + E C 011 i C ˙ 100 = F a C 000 + ( Δ a i κ 2 ) C 100 + g C 011 + F a C 200 , i C ˙ 011 = E C 000 + g C 100 + ( Δ a 2 i κ 2 ) C 011 + F a C 111 + 2 E C 022 , i C ˙ 200 = 2 F a C 100 + 2 ( Δ a i κ 2 ) C 200 + 2 g C 111 , i C ˙ 111 = E C 100 + F a C 011 + 2 g C 200 + ( 2 Δ a 3 i κ 2 ) C 111 + 2 g C 022 i C ˙ 022 = 2 E C 011 + 2 g C 111 + 2 ( Δ a 2 i κ 2 ) C 022
F a C 100 + E C 011 = 0 , F a C 000 + ( Δ a i κ 2 ) C 100 + g C 011 + F a C 200 = 0 , E C 000 + g C 100 + ( Δ a 2 i κ 2 ) C 011 + F a C 111 + 2 E C 022 = 0 , 2 F a C 100 + 2 ( Δ a i κ 2 ) C 200 + 2 g C 111 = 0 , E C 100 + F a C 011 + 2 g C 200 + ( 2 Δ a 3 i κ 2 ) C 111 + 2 g C 022 = 0 , 2 E C 011 + 2 g C 111 + 2 ( Δ a 2 i κ 2 ) C 022 = 0 ,
[ g 1 = λ 2 , E 1 = 1 φ ε ] , [ g 2 = λ 2 , E 2 = 1 φ ε ] ,
ρ ^ t = i [ H ^ e f f , ρ ] + κ a 2 ( n ¯ t h + 1 ) ( 2 a ^ ρ ^ a ^ + 1 2 a ^ a ^ ρ ^ + 1 2 ρ ^ a ^ a ^ ) + κ b 2 ( n ¯ t h + 1 ) ( 2 b ^ ρ ^ b ^ + 1 2 b ^ b ^ ρ ^ + 1 2 ρ ^ b ^ b ^ ) + κ c 2 ( n ¯ t h + 1 ) ( 2 c ^ ρ ^ c ^ + 1 2 c ^ c ^ ρ ^ + 1 2 ρ ^ c ^ c ^ ) + κ a 2 n ¯ t h ( 2 a ^ ρ ^ a ^ + 1 2 a ^ a ^ ρ ^ + 1 2 ρ ^ a ^ a ^ ) + κ b 2 n ¯ t h ( 2 b ^ ρ ^ b ^ + 1 2 b ^ b ^ ρ ^ + 1 2 ρ ^ b ^ b ^ ) + κ c 2 n ¯ t h ( 2 c ^ ρ ^ c ^ + 1 2 c ^ c ^ ρ ^ + 1 2 ρ ^ c ^ c ^ ) ,
g ( 2 ) ( 0 ) = a ^ a ^ a ^   a ^ a ^ a ^ 2 ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.